Next Article in Journal
An Assessment of the Chemical Compatibility of Viton Fluoropolymers and Some Harsh Organic Liquid Mixtures
Next Article in Special Issue
Bioavailability, Sources, and Transfer Behavior of Heavy Metals in Soil–Crop Systems from a High Geological Background Area Impacted by Artisanal Zn Smelting in Guizhou Province, Southwest China
Previous Article in Journal
A Fault Diagnosis Analysis of Afterburner Failure of Aeroengine Based on Fault Tree
Previous Article in Special Issue
A CFD-Based Methodology for Impact Assessment of Industrial Emissions and Evaluation of Mitigation Measures for Regulatory Purposes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effects of Calcination Temperature and Calcination Atmosphere on the Performance of Co3O4 Catalysts for the Catalytic Oxidation of Toluene

College of Environment and Resources, Xiangtan University, Xiangtan 411105, China
*
Author to whom correspondence should be addressed.
Processes 2023, 11(7), 2087; https://doi.org/10.3390/pr11072087
Submission received: 24 February 2023 / Revised: 18 May 2023 / Accepted: 16 June 2023 / Published: 13 July 2023
(This article belongs to the Special Issue Advances in Environmental Pollution and Control Processes)

Abstract

:
A series of Co3O4 catalysts were synthesized and derived from Co-BTC (BTC = 1,3,5-benzenetricarboxylic acid). The effects of different calcination temperatures and calcination atmospheres on the catalytic activity of the materials were investigated. The characteristics of the catalysts were investigated by using various techniques, including X-ray diffraction, N2 adsorption–desorption measurements, scanning electron microscopy, X-ray photoelectron spectroscopy, and H2 temperature-programmed reduction. The findings demonstrated that an increase in calcination temperature caused a higher agglomeration of grains, reduced the specific surface area, and influenced the contents of the active substance Co3+ and surface-adsorbed oxygen of the catalyst. The catalyst pretreated under the N2 atmosphere showed a more uniform particle distribution, better low-temperature reducibility, and the highest catalytic activity. The in situ DRIFTS results indicated that toluene was decomposed successively to benzaldehyde, benzoic acid, bicarbonate, and carbonate species and was eventually broken down into small molecules of CO2 and H2O as the temperature increased.

1. Introduction

Volatile organic compounds (VOCs) are generally defined as the sum of a range of carbon-based organic air pollutants with boiling points ranging from 50 to 260 °C and a Reid vapor pressure greater than 10.3 Pa at a normal temperature (293.15 K) and pressure (101.325 kPa) [1,2]. VOCs have the characteristics of easy volatility, easy diffusion, and easy poisoning and have become major pollutants in the atmosphere [3,4]. Toluene is usually considered to be one of the most typical representatives of VOCs, with a wide range of sources and easy poisoning characteristics in various types of air pollution, such as ozone pollution, fine particulate matter pollution, and secondary aerosol pollution, and it is often an important precursor of direct or indirect harm to the environment and human body, causing widespread concern in society [5,6]. Among the reported methods for toluene removal, catalytic oxidation is considered to be one of the most effective and economical ways to remove toluene, due to its low energy consumption, fast reaction rate, high feasibility, and lack of secondary pollution [7,8].
The choice of catalyst in catalytic oxidation plays a key role in the degradation efficiency of VOCs. The metal catalysts currently used for the degradation of toluene can be divided into noble metal catalysts and non-noble metal oxide catalysts [9,10]. Although noble metal catalysts exhibit outstanding performance for the catalytic removal of VOCs at low temperatures, their susceptible poisoning and high costs have limited their widespread application. Therefore, transition metal oxides have been considered as an alternative for the oxidation of VOCs, due to their relatively low cost, superior reducibility, and anti-poisoning ability [11,12]. Among the non-noble metal oxide catalysts, Co3O4 has the advantages of low cost and being environmentally friendly [13,14]. In related studies, Co3O4 catalysts showed excellent catalytic activity in the degradation process of toluene, and the high activity of Co3O4 catalysts is related to the presence of mobile oxygen in their spinel-type structures, which gives them good low-temperature reduction ability and a large number of oxygen vacancies and adsorbed oxygen content on the surface [15,16]. However, the specific surface area, pore structure, and physical and chemical properties of Co3O4 catalysts also play an important role in the degradation of toluene [17]. Therefore, most of the research is devoted to the synthesis of Co3O4 catalysts with different structures by different preparation methods [13,17]; however, traditional methods are not conducive to controlling the particle size and structure within the catalyst [18]. Therefore, a simple and controllable method is needed to prepare Co3O4 catalysts to achieve high-efficiency degradation of toluene.
In recent years, metal–organic frameworks (MOFs) have attracted widespread interest due to their large specific surface areas, flexible and tunable structures, and versatile properties [19,20]. The nature of the MOFs—namely, that their highly ordered metal nodes are well separated by long chains of organic ligands [21]—also allows the nanometal oxides prepared by direct pyrolysis of MOFs to have highly dispersed metal active centers while still retaining their original pore structure [20]. This provides a very effective method for the design of Co3O4 catalysts. It has been demonstrated that Co3O4 catalysts derived from Co-MOFs showed good degradation of toluene. Juan Lei and his co-workers synthesized Co3O4 catalysts with octahedral, dodecahedral, and rod shapes using three different Co-MOFs; among them, the ZSA-1-Co3O4 catalyst with an octahedral shape showed the best toluene degradation activity, with 90% toluene conversion reached at 249 °C using a gaseous hourly space velocity (GHSV) of 40,000 mL·g−1·h−1 for 1000 ppm of toluene [22].
However, the monometallic catalysts derived from MOFs are generally vulnerable to clustering or sintering during pyrolysis, which leads to larger metallic particles and poorer dispersion, resulting in the reduced catalytic activity of the catalyst [23]. Juan Lei and his colleagues investigated the effects of different calcination temperatures on the catalytic activity of Co3O4 catalysts and found that the catalysts exhibited different physical and chemical properties under different calcination temperatures, with Co3O4 prepared at 350 °C having the largest specific surface area and pore volume and the highest Co3+/Co2+ and Oads/Olatt values, meaning the highest catalytic activity in the catalytic degradation of toluene [24]. Xiaodong Zhang and his co-workers prepared Mn2O3 catalysts in different pyrolysis atmospheres and found that the catalytic activity of the Mn2O3 catalyst pretreated with Ar and then with O2 had the highest Mn3+ and adsorbed oxygen content, and its catalytic activity was superior to that of the Mn2O3 catalysts treated with either Ar or O2 alone [25]. However, most of the existing studies have been limited to investigating the effects of the calcination temperature on the catalysts or the effects of the calcination atmosphere on the catalysts; the synergistic influences of the calcination temperature and atmosphere on the physical and chemical properties of materials have not been comprehensively analyzed.
Herein, we provide a simple method using Co-MOFs as a precursor to obtain honeycomb–spherical structures and Co3O4 catalysts for the degradation of toluene. In order to determine the influence mechanism of the Co3O4 catalysts formed by the calcination temperature and calcination atmosphere on the degradation activity of toluene, the ideal synthetic route for the synthesis of Co3O4 materials using Co-BTC as a precursor was explored to provide a reference for the synthesis process of metal oxides derived from MOFs. Finally, the catalyst was subjected to in situ IR tests under different atmospheric conditions to explore the toluene degradation process and intermediate products.

2. Experimental Section

2.1. Preparation of Catalysts

2.1.1. Synthesis of Co-BTC

Co-BTC was synthesized by a simple solvothermal method. Typically, 0.01 mol of BTC was added to 120 mL of a mixture of DMF, ethanol, and water, with a volume ratio of 8:1:1. After the mixture was completely dissolved, 0.012 mol of Co(NO3)2·6H2O was added to the mixed solution and stirred for 30 min with a magnetic stirrer. Then, the resulting solution was poured into a 200 mL Teflon-lined autoclave and heated at 150 °C for 24 h. After cooling naturally to room temperature, the product was separated using sand core filters. The collected precipitate was added to 30 mL of ethanol and stirred using a magnetic stirrer for 30 min to remove the unreacted chemicals from the material, before being filtered by repeated rinsing with ethanol. The obtained product was dried at 80 °C overnight to obtain the Co-BTC material.

2.1.2. Synthesis of Co3O4 Catalysts

Co-BTC was calcined in air at different temperatures (350 °C, 450 °C, and 550 °C) for 2 h at a heating rate of 2 °C·min−1; the products were named Co3O4-350, Co3O4-450, and Co3O4-550, respectively. Then, the Co-BTC was calcined at 450 °C in N2 for 2 h at a heating rate of 2 °C·min−1 to obtain Co3O4-N2. Additionally, the sample of Co-BTC was first treated at 450 °C in N2 for 1 h and then treated at 350 °C in O2 for 1 h, and the product was named Co3O4-N2-O2. As a comparison, 0.012 mol of the Co(NO3)2·6H2O was treated at 450 °C in N2 for 1 h and then treated at 350 °C in air for 1 h to obtain Co3O4-P.

2.2. Characterization of Materials

The thermal stability of Co-BTC was tested via thermogravimetric analysis (TGA) using an STA-449 F4 Jupiter thermal analyzer (Netzsch, Germany). The weight change in the catalyst was recorded from room temperature to 750 °C under different air and N2 conditions, at a heating rate of 2 °C·min−1. Powder X-ray diffraction (XRD) patterns were recorded on a Bruker D8 Advance powder diffraction system using Cu-Kα radiation in a 2θ range of 5–90°, at a scanning rate of 5°·min−1. The catalysts’ specific surface areas, pore volumes, and pore size distribution were determine using a Micromeritics ASAP 2460 Automated Gas Sorption analyzer. Surface areas were determined using the Brunauer−Emmett−Teller (BET) method; pore volumes and pore size distribution were determined using the Barrett−Joyner−Halenda (BJH) method. A ZEISS Sigma 300 scanning electron microscope (SEM) (Oberkochen, Germany) was operated at an accelerating voltage of 15 kV. The catalyst was deposited on a clean silicon wafer, mounted on the aluminum holder using carbon tape, and sputter-coated with gold to minimize sample charging. X-ray photoelectron spectroscopy (XPS) analysis was performed using a Thermo Fisher Scientific spectrophotometer with Al Kα (1486.6 eV) as an X-ray excitation source. Survey scans and narrow scans were carried out at pass energies of 100 eV and 30 eV, respectively. C1 s was used as a reference and assigned a binding energy of 284.8 eV. The Micromeritics AutoChem II 2920 chemical adsorption apparatus was used for a hydrogen temperature-programmed reduction (H2-TPR) experiment. Before the H2-TPR experiment, 0.1 g of catalyst was pretreated under an Ar atmosphere at 150 °C for 1 h. The temperature was then cooled to 50 °C and the reduction process began, proceeding in a temperature range of 50–800 °C in 10% vol. H2/Ar (30 mL·min−1). In situ diffusion reflectance infrared Fourier-transform spectrophotometry (DRIFTS; Nicolet is 50) was employed to investigate the reaction process of toluene oxidation over Co3O4-N2-O2. The catalyst was squeezed into a thin slice in the DRIFT cell and tested in situ using absorbance; it was then activated to remove surface impurities under Ar for 1 h at 200 °C, and then cooled to room temperature. Toluene (200 ppm) was introduced until the peaks remained unchanged. Subsequently, Ar or air was used to record changes in IR peaks at different temperatures under different gas atmospheres.

2.3. Catalytic Testing

The catalytic oxidation of toluene was carried out in a fixed-bed flow reactor at atmospheric pressure. First, 100 mg of catalyst was packed into a quartz tube 10 cm in length, with an inner diameter of 2 cm; the ends were held in place by quartz fiber, and the reaction temperature was controlled using a tubular furnace (KTF-0.5-06). Toluene was injected from a single-channel syringe pump (SPLab01) into the gasification chamber, and the gasified toluene was carried through 100 mL·min−1 air into a quartz tube (the gas flow was controlled using a gas flowmeter). For toluene, the reacted gas was detected using gas chromatography (GC112A), and the detection conditions were as follows: high-purity N2 was used as a carrier gas; a flame ionization detector (FID) was used; the column temperature was 150 °C; the gasification chamber temperature was 250 °C; the detector temperature was 250 °C; and the chromatographic column was a capillary column. For CO2, the reacted gas was detected using a multifunctional gas analyzer (GT-2000). Toluene’s conversion (η) and mineralization (μ) were calculated according to the following equations:
η = C 0 C i C 0 × 100 %
μ = C C O 2 , o u t C C O 2 , i n × 100 %
where C0 (ppm) and Ci (ppm) are the toluene concentrations of the inlet and outlet gases, respectively, while C C O 2 , i n (ppm) and C C O 2 , o u t (ppm) represent the concentration of carbon dioxide at the time of complete oxidation of toluene and the concentration of carbon dioxide at the outlet after the reaction, respectively.
Reaction rates over catalysts and apparent activation energies (Ea) were calculated for toluene conversions lower than 20%, in line with the practice previously described in the literature [24].

3. Results and Discussion

3.1. Structural Analysis

The XRD patterns of all catalysts are presented in Figure 1. It can be seen that the air-calcined Co3O4 catalysts in the 2θ range of 5–90° displayed similar patterns, which corresponded to the standard Co3O4 sample (Co3O4-PDF#74-2120). The characteristic diffraction peaks of Co3O4 at 19°, 31.3°, 36.8°, 44.8°, 59.4°, and 65.2° corresponded to the (111), (220), (311), (400), (511), and (440) crystal faces, respectively. This showed that the air-calcined Co-BTC ended up with a Co3O4 crystal-phase structure. The Co3O4-N2 nanomaterial exhibited faint diffraction peaks at 2θ of 36.6°, 42.6°, and 61.8°, which corresponded to the (111), (200), and (220) lattice planes, respectively, of pure CoO (CoO-PDF#75-0533). It was found that the Co in the Co3O4-N2 catalyst existed mainly in the form of CoO, due to the fact that, in the absence of oxygen, Co mainly combines with oxygen bonds to form CoO. From Figure 1 and Tables S1 and S2, it can clearly be seen that the peak intensity of Co3O4 was enhanced and its width narrowed with increasing air calcination temperature, indicating that with the increase in calcination temperature, O2 oxidized the metal clusters in the Co-BTC skeleton more fully, resulting in better crystallization of Co3O4 and the formation of metal oxide clusters with larger particles. By comparing the diffraction peaks of the metal oxides formed in different calcination atmospheres, it was found that the oxides calcined with N2 had lower peak intensities and larger peak widths, especially for the Co3O4-N2 catalyst, indicating that the Co3O4-N2 catalyst had a smaller particle size.
The N2 adsorption−desorption isotherms and BJH pore size distributions of the Co3O4 catalysts are shown in Figure S2. The catalysts all exhibited type IV isotherms with H3-type hysteresis loops, indicating that the catalysts had similar mesoporous structures [26,27]. Table 1 shows the specific surface area, average pore size, and pore volume data for the different catalysts. It can be seen that the influence of the calcination temperature on the pore structure of the material was significant. With the increase in temperature, the specific surface area, average pore size, and pore volume of the material all decreased as a result of agglomeration and expansion between crystals at high calcination temperatures (Figure 2). In addition, the calcination atmosphere also affected the pore structure of the material. It was found that the oxides calcined under an N2 atmosphere had a larger specific surface area. Co3O4-N2 had the largest specific surface area (161.4 m2·g−1), and this was attributed to the carbonization of the organic ligands in Co-BTC under the N2 atmosphere, which enabled the material to maintain its original skeletal structure. It was also found that Co3O4-N2-O2 had the largest average pore size, which facilitated the transfer process between reactants and products.
The microstructures of the catalysts are presented in Figure 2. It can be seen that all catalysts exhibited a honeycomb-like spherical structure. From Figure 2a–c, it can also be seen that the pore structure inside the Co3O4-350 catalyst spheres was composed of a number of aggregated grains. As the air calcination temperature was increased, the surface of these grains gradually became smooth, and the particle size also showed signs of expansion, indicating that sintering and agglomeration between the grains occurred at high temperatures, resulting in a decreasing specific surface area of the catalyst, which was consistent with the results of the XRD and BET characterizations. The Co3O4-N2-O2 catalyst spheres showed no obvious aggregation between the grains and exhibited a highly discrete and uniform distribution; this indicated that the catalyst exhibited more roughness and possessed a larger specific surface area, facilitating the exposure of the Co active sites in the catalyst, and thereby increasing the chance of contact with toluene and improving the catalytic oxidation activity of the catalysts.

3.2. Surface Composition and Reduction Behavior

To obtain further information about elemental chemical states and surface oxygen species, the Co 2p3/2 and O 1s XPS spectra of the samples were obtained, as shown in Figure 3a,b, respectively. The Co 2p3/2 XPS peak could be fitted to two component peaks and two satellite peaks, with the main peaks at binding energies of 779.5 eV and 781.29 eV corresponding to the Co3+ and Co2+ species, and the main peaks at binding energies of 785.68 eV and 789.02 eV corresponding to the Co3+ and Co2+ satellite peaks [28,29]. The Co3+ and Co2+ contents and the corresponding Co3+/Co2+ ratios were calculated from the peak areas, and the data are presented in Table 2 and Table 3. The ratio of Co3+/Co2+ increased in the order of Co3O4-N2 < Co3O4-550 < Co3O4-450 < Co3O4-350 ≈ Co3O4-N2-O2, where Co3O4-350 and Co3O4-N2-O2 exhibited the highest values of Co3+/Co2+, indicating the presence of abundant Co3+ species in Co3O4-350 and Co3O4-N2-O2. Previous studies have shown that Co3+ usually acts as the active center of toluene and can effectively promote the degradation of toluene [30].
Figure 3b shows the O 1s XPS spectra of the samples, which could be divided into two peaks at binding energies of 529.36 eV and 531.03 eV, respectively. This could be ascribed to lattice oxygen (Olatt) and adsorbed oxygen species on oxygen vacancies (Oads) [31], where the Co3O4-N2 catalyst shifted the position of the oxygen species peak, so that the peaks with binding energies at 529.66 eV and 531.54 eV represented lattice oxygen species and adsorption oxygen, respectively. The Oads/Olatt molar ratios were determined based on the area-fitted curve peaks of the O 1s XPS spectra; these are listed in Table 2 and Table 3. It can be seen that the ratio of Oads/Olatt decreased as the calcination temperature of the material increased, indicating that high-temperature calcination was not conducive to the accumulation of adsorbed oxygen in the material. Previous studies have shown that surface-adsorbed oxygen (Oads) is related to surface defects and that higher contents of Oads facilitate the uptake and activation of gas-phase oxygen molecules by the catalyst, promoting the degradation of pollutants [32,33]. In addition, it is worth noting that Co3O4-N2 had the highest Oads/Olatt molar ratios. However, from Figure S3 and Table 3, it can also be seen that the surface C, O, and Co contents of the Co3O4-350 and Co3O4-N2-O2 catalysts were essentially the same, with the highest elemental content being that of O, accounting for about 50% of the total. In contrast, the Co3O4-N2 catalyst had the largest amount of elemental C but only 23% of the total amount of O. This was because the Co3O4-N2 catalyst was prepared in the absence of oxygen, so most of the organic matter in the ligand was stored as carbon inside the catalyst, and the metal Co inside the material could not be fully oxidized, resulting in a lack of sufficient O in the catalyst itself, so that the Co3O4-N2 catalyst had the highest Oads/Olatt molar ratios but the lowest total amount of Oads compared with other catalysts. This phenomenon can also be seen in Figure S1.
The reducibility of a sample was highly related to its catalytic performance. Figure 4 illustrates the H2-TPR profiles of all synthesized catalysts. Three distinct reduction peaks were found for all catalysts, in which the low-temperature-region reduction peak indicated the reduction of Co3+ to Co2+, the medium-temperature-region reduction peak indicated the reduction of Co2+ to Co+, and the high-temperature-region reduction peak indicated the reduction of Co2+ to Co0 [34]. From Figure 4a, it can be seen that the reduction peak of the catalyst gradually moved towards the high-temperature region as the calcination temperature increased, indicating that the reduction effect of the material deteriorated under high-temperature calcination conditions; this may be related to the particle size of the catalyst, as the XRD results indicated that the particle size of the catalyst became larger as the calcination temperature increased. It has been reported in the literature that the reduction peak tends to move towards the high-temperature region as the particle size increases [35]. Figure 4b shows the reduction peaks of the catalysts prepared in different calcination atmospheres. A comparison reveals that Co3O4-N-O2 had the lowest reduction temperature. This finding may be related to the specific surface area of the material and the abundance of Co3+ active centers, i.e., the high specific surface area of the material made it easier to expose a large number of Co3+ active sites on the external surface. Compared with Co3O4-N2-O2, the reduction peak of Co3O4-N2 was very weak, because the Co in the Co3O4-N2 catalyst existed mainly in the form of CoO and, thus, lacked a sufficient number of Co3+ active sites.

3.3. Catalytic Performance

Figure 5a,b show the temperature dependence of the catalytic efficiency in toluene conversion for all as-prepared catalysts under the conditions of toluene concentration = 1000 ppm and WHSV = 60,000 mL·g−1·h−1. It can be seen from the graph that no catalyst showed any significant toluene degradation below 210 °C; subsequently, as the temperature increased, the catalysts’ active sites were gradually activated and the toluene degradation rate rapidly increased. From Figure 5a, it can be seen that increasing the calcination temperature of the catalyst did not promote catalytic activity, because this increase in calcination temperature also caused the particle size of the catalyst to gradually increase, resulting in a decreased specific surface area of the catalyst, which was not conducive to the exposure of the active sites. From Figure 5b, it can be seen that the different calcination atmospheres of the materials also had an effect on the degradation of toluene. The Co3O4-N2-O2 catalyst exhibited the best catalytic effect, while the Co3O4-N2 catalyst had the poorest, i.e., the worst degradation of toluene. This was because the catalytic activity of the catalyst was not only related to the specific surface area of the material—the Co3+ active sites and surface-adsorbed oxygen content were also important influencing factors. The degradation activity of the catalyst for toluene decreased as follows: Co3O4-N2-O2 > Co3O4-350 > Co3O4-450 > Co3O4-N2 > Co3O4-550. The best toluene degradation rates were produced by Co3O4-N2-O2. The T10, T50, and T90 temperatures required for 10%, 50%, and 90% toluene conversion were 221 °C, 234 °C, and 245 °C, respectively. In combination with the above characterization, these findings demonstrate that the catalytic activity of the catalysts was determined by a combination of the specific surface area, the morphology, the contents of Co3+ active sites and surface-adsorbed oxygen, and the redox properties of the material. In short, we found that a large specific surface area, homogeneous grain distribution, high Co3+ and surface-adsorbed oxygen contents, and excellent redox properties promoted the catalytic activity of the material.
The apparent activation energy (Ea) of the catalysts was calculated by applying the linear fitting of Arrhenius plots at a conversion rate of less than 20% (as shown in Figure 5c). The calculated results are listed in Table S3 in the following order: Co3O4-N2-O2 (39.65 kJ·mol−1) < Co3O4-350 (42.06 kJ·mol−1) < Co3O4-450 (47.84 kJ·mol−1) < Co3O4-N2 (67.27 kJ·mol−1) < Co3O4-550 (70.43 kJ·mol−1). The lowest Ea value corresponds to the highest possibility of toluene decomposition. These results were consistent with those of toluene conversion.
In order to explore the advantages of Co3O4 catalysts derived from Co-MOF in terms of catalytic activity and mineralization rate, we carried out a test in which Co3O4-N2-O2 (with the best catalytic activity) was used as the catalyst. The toluene conversion rates and CO2 yields were measured for both the Co3O4-N2-O2 and Co3O4-P catalysts. As can be seen in Figure 5d,e, we found that the Co3O4-N2-O2 catalyst achieved obviously higher conversion rates at lower temperatures than the Co3O4-P catalyst, in terms of both toluene degradation efficiency and mineralization efficiency. The toluene degradation efficiencies of Co3O4-N2-O2 catalysts at 50% and 90% were 234 °C and 245 °C, respectively, and these values were broadly consistent with the temperatures corresponding to toluene mineralization efficiencies at 50% and 90% (236 °C and 246 °C, respectively). This showed that the Co3O4-N2-O2 catalyst could adequately convert toluene molecules to CO2 and H2O at the appropriate temperature, thereby reducing the emission of intermediate products during the reaction.
Figure 5f shows results of a continuous test of the catalytic oxidation of toluene over the Co3O4-N2-O2 catalyst. The test results demonstrated that the toluene degradation efficiency of the catalyst remained above 99% for 50 h at 260 °C, without any significant decline, indicating that the catalytic activity of the Co3O4-N2-O2 catalyst does not change over time and exhibits good catalytic stability when operated for a long period at an appropriate temperature, suggesting great potential for its usage in practical applications.

3.4. Mechanism for Degradation of Toluene over Co3O4-N2-O2

To further investigate the possible reaction pathway of toluene oxidation on the Co3O4-N2-O2 catalyst and any intermediates in the degradation of toluene, in situ DRIFT spectra of the catalyst at different temperatures under Ar and air conditions were obtained, as shown in Figure 6. It was found that, after a period of introducing toluene at room temperature, characteristic peaks of toluene appeared: at 3091 and 3035 cm−1, attributed to unsaturated C-H bonds on the benzene ring; and at 2932 cm−1, attributed to saturated C-H bonds on toluene’s methyl groups. In addition. 1585 and 1495 cm−1 were found to be the benzene ring skeleton vibration peaks of toluene [36,37,38], indicating that toluene was adsorbed on the surface of the catalyst. Figure 6a shows that the characteristic peak of toluene did not change significantly with increasing temperature under an Ar atmosphere, meaning that the catalyst had little degradation effect on toluene under an Ar atmosphere. Conversely, under air-supply conditions, as shown in Figure 6b, the characteristic peak of toluene disappeared with the increase in temperature. At 140 °C, new bands appeared at 2840 and 2721 cm−1, which could possibly be attributed to adsorbed benzaldehyde [39]; in addition, bicarbonate species (1356 cm−1) also started to appear [40]. New characteristic peaks also appeared at 1563 and 1340 cm−1 when the temperature reached 200 °C; these were attributed to benzoic acid and carbonate species, respectively [41,42]. As the temperature increased, benzaldehyde was further oxidized to benzoic acid, and bicarbonate species were gradually oxidized to carbonate species. The characteristic peaks of 1431 cm−1 for carboxylate species such as benzoic acid, and of 2340 and 2360 cm−1 for CO2, became progressively larger as the temperature increased under an air atmosphere [38], meaning that more and more benzoic acid and CO2 were generated and accumulated on the catalyst’s surface as the temperature increased.
The above analysis clearly demonstrates that the toluene molecules in the air were adsorbed on the surface of the catalyst, and that the catalytic oxidation effect of the lattice oxygen of the catalyst on toluene adsorbed on the catalyst’s surface was not significant. The degradation process of toluene molecules is shown in Figure 7. The toluene molecules were first adsorbed on the catalyst’s surface and then gradually degraded by the interaction with the adsorbed oxygen. With increasing temperature, the toluene molecules were gradually oxidized into benzaldehyde, benzoic acid, and bicarbonate and carbonate species, and finally decomposed into small molecules of CO2 and H2O [42].

4. Conclusions

Different Co3O4 catalysts were prepared under different calcination temperatures and calcination atmospheres, and their degradation effects on toluene were compared. It was found that increasing the air calcination temperature intensified the agglomeration of the internal grains of the catalyst and reduced the contents of Co3+ active sites and surface-adsorbed oxygen; this was not conducive to the improvement of the catalytic activity of the material. However, pretreatment with N2 significantly improved the internal structure of the catalyst, making the internal grains of the catalyst more uniformly distributed; this promoted the exposure of active sites and improved the degradation efficiency of toluene. The Co3O4-N2-O2 catalyst exhibited superior toluene degradation activity and better toluene mineralization rates compared with Co3O4-P; it also exhibited stable catalytic activity, indicating its great potential for practical applications. Finally, the degradation process of toluene on the surface of Co3O4-N2-O2 catalyst was investigated; it was found that benzaldehyde, benzoic acid, and bicarbonate and carbonate species are intermediate products in the degradation process of toluene, which are ultimately decomposed into small molecules of CO2 and H2O with increased temperature.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/pr11072087/s1. Figure S1: Thermal weight curve of Co-BTC; Figure S2: N2 adsorption and desorption isotherms of samples (a) different calcination atmosphere and (b) calcination temperature in air, and (c) pore-size distributions of all samples. Figure S3: XPS full spectra of Co3O4 catalysts. Table S1: Peak height of the XRD diffraction peaks of the catalysts. Table S2: FWHM of the XRD diffraction peaks of the catalysts. Table S3: Catalytic activities of toluene and apparent activation energies (Ea) of all catalysts.

Author Contributions

S.J.: Data curation, Software, Writing-original draft. Z.Y.: Conceptualization, Methodology, Writing—review & editing. N.T.: Supervision. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in the article and its supplementary material.

Acknowledgments

We wish to thank College of Environment and Resources, Xiangtan University for providing a congenial work environment and state-of-the-art research facilities. We also thank Zhimin You for his valuable support, comments, and suggestions.

Conflicts of Interest

The authors declare no conflict of interest.

Declaration of Competing Interest

The authors declare that they have no known competing financial interest or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Kamal, M.S.; Razzak, S.A.; Hossain, M.M. Catalytic oxidation of volatile organic compounds (VOCs)–A review. Atmos. Environ. 2016, 140, 117–134. [Google Scholar] [CrossRef]
  2. Dai, C.; Zhou, Y.; Peng, H.; Huang, S.; Qin, P.; Zhang, J.; Yang, Y.; Luo, L.; Zhang, X. Current progress in remediation of chlorinated volatile organic compounds: A review. J. Ind. Eng. Chem. 2018, 62, 106–119. [Google Scholar] [CrossRef]
  3. Li, J.; Lu, R.; Dou, B.; Ma, C.; Hu, Q.; Liang, Y.; Wu, F.; Qiao, S.; Hao, Z. Porous Graphitized Carbon for Adsorptive Removal of Benzene and the Electrothermal Regeneration. Environ. Sci. Technol. 2012, 46, 12648–12654. [Google Scholar] [CrossRef] [Green Version]
  4. Jiang, Z.; Feng, X.; Deng, J.; He, C.; Douthwaite, M.; Yu, Y.; Liu, J.; Hao, Z.; Zhao, Z. Atomic-Scale Insights into the Low-Temperature Oxidation of Methanol over a Single-Atom Pt1-Co3O4 Catalyst. Adv. Funct. Mater. 2019, 29, 1902041. [Google Scholar] [CrossRef]
  5. Zhang, Q.; Yuan, B.; Shao, M.; Wang, X.; Lu, S.; Lu, K.; Wang, M.; Chen, L.; Chang, C.C.; Liu, S. Variations of ground-level O-3 and its precursors in Beijing in summertime between 2005 and 2011. Atmos. Chem. Phys. 2014, 14, 6089–6101. [Google Scholar] [CrossRef] [Green Version]
  6. Yang, C.; Miao, G.; Pi, Y.; Xia, Q.; Wu, J.; Li, Z.; Xiao, J. Abatement of various types of VOCs by adsorption/catalytic oxidation: A review. Chem. Eng. J. 2019, 370, 1128–1153. [Google Scholar] [CrossRef]
  7. Jiang, Z.; He, C.; Dummer, N.F.; Shi, J.; Tian, M.; Ma, C.; Hao, Z.; Taylor, S.H.; Ma, M.; Shen, Z. Insight into the efficient oxidation of methyl-ethyl-ketone over hierarchically micro-mesostructured Pt/K-(Al)SiO2 nanorod catalysts: Structure-activity relationships and mechanism. Appl. Catal. B Environ. 2018, 226, 220–233. [Google Scholar] [CrossRef] [Green Version]
  8. Tian, M.; He, C.; Yu, Y.; Pan, H.; Smith, L.; Jiang, Z.; Gao, N.; Jian, Y.; Hao, Z.; Zhu, Q. Catalytic oxidation of 1,2-dichloroethane over three-dimensional ordered meso-macroporous Co3O4/La0.7Sr0.3Fe0.5Co0.5O3: Destruction route and mechanism. Appl. Catal. A Gen. 2018, 553, 1–14. [Google Scholar] [CrossRef] [Green Version]
  9. Zhang, X.; Song, L.; Bi, F.; Zhang, D.; Wang, Y.; Cui, L. Catalytic oxidation of toluene using a facile synthesized Ag nanoparticle supported on UiO-66 derivative. J. Colloid Interface Sci. 2020, 571, 38–47. [Google Scholar] [CrossRef] [PubMed]
  10. Mo, S.; Zhang, Q.; Li, J.; Sun, Y.; Ren, Q.; Zou, S.; Zhang, Q.; Lu, J.; Fu, M.; Mo, D.; et al. Highly efficient mesoporous MnO2 catalysts for the total toluene oxidation: Oxygen-Vacancy defect engineering and involved intermediates using in situ DRIFTS. Appl. Catal. B Environ. 2020, 264, 118464. [Google Scholar] [CrossRef]
  11. He, C.; Cheng, J.; Zhang, X.; Douthwaite, M.; Pattisson, S.; Hao, Z. Recent Advances in the Catalytic Oxidation of Volatile Organic Compounds: A Review Based on Pollutant Sorts and Sources. Chem. Rev. 2019, 119, 4471–4568. [Google Scholar] [CrossRef] [PubMed]
  12. Fang, Y.; Guo, Y. Copper-based non-precious metal heterogeneous catalysts for environmental remediation. Chin. J. Catal. 2018, 39, 566–582. [Google Scholar] [CrossRef]
  13. Ren, Q.; Feng, Z.; Mo, S.; Huang, C.; Li, S.; Zhang, W.; Chen, L.; Fu, M.; Wu, J.; Ye, D. 1D-Co3O4, 2D-Co3O4, 3D-Co3O4 for catalytic oxidation of toluene. Catal. Today 2019, 332, 160–167. [Google Scholar] [CrossRef]
  14. Zhang, Q.; Mo, S.; Chen, B.; Zhang, W.; Huang, C.; Ye, D. Hierarchical Co3O4 nanostructures in-situ grown on 3D nickel foam towards toluene oxidation. Mol. Catal. 2018, 454, 12–20. [Google Scholar] [CrossRef]
  15. Kołodziej, A.; Łojewska, J.; Tyczkowski, J.; Jodłowski, P.; Redzynia, W.; Iwaniszyn, M.; Zapotoczny, S.; Kuśtrowski, P. Coupled engineering and chemical approach to the design of a catalytic structured reactor for combustion of VOCs: Cobalt oxide catalyst on knitted wire gauzes. Chem. Eng. J. 2012, 200–202, 329–337. [Google Scholar] [CrossRef]
  16. Liu, Q.; Wang, L.-C.; Chen, M.; Cao, Y.; He, H.-Y.; Fan, K.-N. Dry citrate-precursor synthesized nanocrystalline cobalt oxide as highly active catalyst for total oxidation of propane. J. Catal. 2009, 263, 104–113. [Google Scholar] [CrossRef]
  17. Ren, Q.; Mo, S.; Peng, R.; Feng, Z.; Zhang, M.; Chen, L.; Fu, M.; Wu, J.; Ye, D. Controllable synthesis of 3D hierarchical Co3O4 nanocatalysts with various morphologies for the catalytic oxidation of toluene. J. Mater. Chem. A 2018, 6, 498–509. [Google Scholar] [CrossRef]
  18. Qian, S.; Wang, C.; Liu, W.; Zhu, Y.; Yao, W.; Lu, X. An enhanced CdS/TiO2 photocatalyst with high stability and activity: Effect of mesoporous substrate and bifunctional linking molecule. J. Mater. Chem. 2011, 21, 4945–4952. [Google Scholar] [CrossRef]
  19. Liu, B.; Vellingiri, K.; Jo, S.-H.; Kumar, P.; Ok, Y.S.; Kim, K.-H. Recent advances in controlled modification of the size and morphology of metal-organic frameworks. Nano Res. 2018, 11, 4441–4467. [Google Scholar] [CrossRef]
  20. He, H.; Li, R.; Yang, Z.; Chai, L.; Jin, L.; Alhassan, S.I.; Ren, L.; Wang, H.; Huang, L. Preparation of MOFs and MOFs derived materials and their catalytic application in air pollution: A review. Catal. Today 2021, 375, 10–29. [Google Scholar] [CrossRef]
  21. Liu, H.; Zhang, S.; Liu, Y.; Yang, Z.; Feng, X.; Lu, X.; Huo, F. Well-Dispersed and Size-Controlled Supported Metal Oxide Nanoparticles Derived from MOF Composites and Further Application in Catalysis. Small 2015, 11, 3130–3134. [Google Scholar] [CrossRef] [PubMed]
  22. Lei, J.; Wang, S.; Li, J. Mesoporous Co3O4 derived from Co-MOFs with different morphologies and ligands for toluene catalytic oxidation. Chem. Eng. Sci. 2020, 220, 115654. [Google Scholar] [CrossRef]
  23. Sager, T.M.; Porter, D.W.; Robinson, V.A.; Lindsley, W.G.; Schwegler-Berry, D.E.; Castranova, V. Improved method to disperse nanoparticles for in vitro and in vivo investigation of toxicity. Nanotoxicology 2007, 1, 118–129. [Google Scholar] [CrossRef]
  24. Lei, J.; Wang, S.; Li, J. Mesoporous Co3O4 Derived from Facile Calcination of Octahedral Co-MOFs for Toluene Catalytic Oxidation. Ind. Eng. Chem. Res. 2020, 59, 5583–5590. [Google Scholar] [CrossRef]
  25. Zhang, X.; Lv, X.; Bi, F.; Lu, G.; Wang, Y. Highly efficient Mn2O3 catalysts derived from Mn-MOFs for toluene oxidation: The influence of MOFs precursors. Mol. Catal. 2020, 482, 110701. [Google Scholar] [CrossRef]
  26. Du, Y.; Meng, Q.; Wang, J.; Yan, J.; Fan, H.; Liu, Y.; Dai, H. Three-dimensional mesoporous manganese oxides and cobalt oxides: High-efficiency catalysts for the removal of toluene and carbon monoxide. Microporous Mesoporous Mater. 2012, 162, 199–206. [Google Scholar] [CrossRef]
  27. Feng, X.; Guo, J.; Wen, X.; Xu, M.; Chu, Y.; Yuan, S. Enhancing performance of Co/CeO2 catalyst by Sr doping for catalytic combustion of toluene. Appl. Surf. Sci. 2018, 445, 145–153. [Google Scholar] [CrossRef]
  28. Mo, S.; Li, S.; Li, J.; Deng, Y.; Peng, S.; Chen, J.; Chen, Y. Rich surface Co(iii) ions-enhanced Co nanocatalyst benzene/toluene oxidation performance derived from Co(II)Co(III) layered double hydroxide. Nanoscale 2016, 8, 15763–15773. [Google Scholar] [CrossRef]
  29. Xu, L.; Jiang, Q.; Xiao, Z.; Li, X.; Huo, J.; Wang, S.; Dai, L. Plasma-Engraved Co3 O4 Nanosheets with Oxygen Vacancies and High Surface Area for the Oxygen Evolution Reaction. Angew. Chem. 2016, 55, 5277–5281. [Google Scholar] [CrossRef]
  30. Wang, X.; Liu, Y.; Zhang, T.; Luo, Y.; Lan, Z.; Zhang, K.; Zuo, J.; Jiang, L.; Wang, R. Geometrical-Site-Dependent Catalytic Activity of Ordered Mesoporous Co-Based Spinel for Benzene Oxidation: In Situ DRIFTS Study Coupled with Raman and XAFS Spectroscopy. ACS Catal. 2017, 7, 1626–1636. [Google Scholar] [CrossRef]
  31. Mo, S.; Li, S.; Li, W.; Li, J.; Chen, J.; Chen, Y. Excellent low temperature performance for total benzene oxidation over mesoporous CoMnAl composited oxides from hydrotalcites. J. Mater. Chem. A 2016, 4, 8113–8122. [Google Scholar] [CrossRef]
  32. Todorova, S.; Kadinov, G.; Tenchev, K.; Kalvachev, Y.; Kostov-Kytin, V. Particle size and support effects on the complete benzene oxidation by Co and Co–Pt catalysts. J. Mater. Sci. 2007, 42, 3315–3320. [Google Scholar] [CrossRef]
  33. Hu, Z.; Liu, X.; Meng, D.; Guo, Y.; Guo, Y.; Lu, G. Effect of Ceria Crystal Plane on the Physicochemical and Catalytic Properties of Pd/Ceria for CO and Propane Oxidation. ACS Catal. 2016, 6, 2265–2279. [Google Scholar] [CrossRef]
  34. Zhao, S.; Hu, F.; Li, J. Hierarchical Core–Shell Al2O3@Pd-CoAlO Microspheres for Low-Temperature Toluene Combustion. ACS Catal. 2016, 6, 3433–3441. [Google Scholar] [CrossRef]
  35. Li, J.-B.; Jiang, Z.-Q.; Qian, K.; Huang, W.-X. Effect of Calcination Temperature on Surface Oxygen Vacancies and Catalytic Performance Towards CO Oxidation of Co3O4 Nanoparticles Supported on SiO2. Chin. J. Chem. Phys. 2012, 25, 103–109. [Google Scholar] [CrossRef]
  36. Sun, H.; Liu, Z.; Chen, S.; Quan, X. The role of lattice oxygen on the activity and selectivity of the OMS-2 catalyst for the total oxidation of toluene. Chem. Eng. J. 2015, 270, 58–65. [Google Scholar] [CrossRef]
  37. Besselmann, S.; Löffler, E.; Muhler, M. On the role of monomeric vanadyl species in toluene adsorption and oxidation on V2O5/TiO2 catalysts: A Raman and in situ DRIFTS study. J. Mol. Catal. A Chem. 2000, 162, 401–411. [Google Scholar] [CrossRef]
  38. Lei, J.; Wang, S.; Li, J.; Xu, Y.; Li, S. Different effect of Y (Y = Cu, Mn, Fe, Ni) doping on Co3O4 derived from Co-MOF for toluene catalytic destruction. Chem. Eng. Sci. 2022, 251, 117436. [Google Scholar] [CrossRef]
  39. Jia, Z.; Wang, X.; Thevenet, F.; Rousseau, A. Dynamic probing of plasma-catalytic surface processes: Oxidation of toluene on CeO2. Plasma Process. Polym. 2017, 14, 1600114. [Google Scholar] [CrossRef] [Green Version]
  40. Araña, J.; González Díaz, O.; Miranda Saracho, M.; Doña Rodríguez, J.M.; Herrera Melián, J.A.; Pérez Peña, J. Photocatalytic degradation of formic acid using Fe/TiO2 catalysts: The role of Fe3+/Fe2+ ions in the degradation mechanism. Appl. Catal. B Environ. 2001, 32, 49–61. [Google Scholar] [CrossRef]
  41. Chen, X.; Chen, X.; Yu, E.; Cai, S.; Jia, H.; Chen, J.; Liang, P. In situ pyrolysis of Ce-MOF to prepare CeO2 catalyst with obviously improved catalytic performance for toluene combustion. Chem. Eng. J. 2018, 344, 469–479. [Google Scholar] [CrossRef]
  42. Du, J.; Qu, Z.; Dong, C.; Song, L.; Qin, Y.; Huang, N. Low-temperature abatement of toluene over Mn-Ce oxides catalysts synthesized by a modified hydrothermal approach. Appl. Surf. Sci. 2018, 433, 1025–1035. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of Co3O4 catalysts.
Figure 1. XRD patterns of Co3O4 catalysts.
Processes 11 02087 g001
Figure 2. SEM images with different magnification (4 μm and 200 nm) of (a-1,a-2) Co3O4-550, (b-1,b-2) Co3O4-450, (c-1,c-2) Co3O4-350, (d-1,d-2) Co3O4-N2-O2, and (e-1,e-2) Co3O4-N2.
Figure 2. SEM images with different magnification (4 μm and 200 nm) of (a-1,a-2) Co3O4-550, (b-1,b-2) Co3O4-450, (c-1,c-2) Co3O4-350, (d-1,d-2) Co3O4-N2-O2, and (e-1,e-2) Co3O4-N2.
Processes 11 02087 g002
Figure 3. XPS spectra of Co3O4 catalysts: (a) Co 2p3/2; (b) O 1s.
Figure 3. XPS spectra of Co3O4 catalysts: (a) Co 2p3/2; (b) O 1s.
Processes 11 02087 g003
Figure 4. H2-TPR spectra of Co3O4 catalysts under (a) different calcination temperatures and (b) different calcination atmospheres.
Figure 4. H2-TPR spectra of Co3O4 catalysts under (a) different calcination temperatures and (b) different calcination atmospheres.
Processes 11 02087 g004
Figure 5. Toluene degradation rates of Co3O4 catalysts under (a) different calcination temperatures; and (b) different calcination atmospheres (0.1 g catalyst; 1000 ppm toluene; WHSV = 60,000 mL·g−1·h−1); (c) Arrhenius plots; (d,e) comparison of (d) toluene catalytic performance and (e) CO2 selectivity of Co3O4-N2-O2 and Co3O4-P (0.1 g catalyst; 1000 ppm toluene; WHSV = 60,000 mL·g−1·h−1); (f) Co3O4-N2-O2 on-stream stability test of continuous catalytic process (0.1 g catalyst; 1000 ppm toluene; WHSV = 60,000 mL·g−1·h−1; T = 260 °C).
Figure 5. Toluene degradation rates of Co3O4 catalysts under (a) different calcination temperatures; and (b) different calcination atmospheres (0.1 g catalyst; 1000 ppm toluene; WHSV = 60,000 mL·g−1·h−1); (c) Arrhenius plots; (d,e) comparison of (d) toluene catalytic performance and (e) CO2 selectivity of Co3O4-N2-O2 and Co3O4-P (0.1 g catalyst; 1000 ppm toluene; WHSV = 60,000 mL·g−1·h−1); (f) Co3O4-N2-O2 on-stream stability test of continuous catalytic process (0.1 g catalyst; 1000 ppm toluene; WHSV = 60,000 mL·g−1·h−1; T = 260 °C).
Processes 11 02087 g005
Figure 6. In situ DRIFT spectra of toluene oxidation of Co3O4-N2-O2 in Ar (a) and air (b).
Figure 6. In situ DRIFT spectra of toluene oxidation of Co3O4-N2-O2 in Ar (a) and air (b).
Processes 11 02087 g006
Figure 7. Degradation process of toluene by Co3O4-N2-O2 catalyst.
Figure 7. Degradation process of toluene by Co3O4-N2-O2 catalyst.
Processes 11 02087 g007
Table 1. Specific surface area and pore structure data of Co3O4 catalysts.
Table 1. Specific surface area and pore structure data of Co3O4 catalysts.
SamplesBET Surface Area
(m2·g−1)
Average Pore Size
(nm)
Pore Volume
(cm3·g−1)
Co3O4-5508.74.90.01
Co3O4-45016.89.40.04
Co3O4-35040.217.90.18
Co3O4-N2-O257.221.250.3
Co3O4-N2161.411.510.46
Table 2. XPS surface element analysis of catalysts prepared at different calcination temperatures.
Table 2. XPS surface element analysis of catalysts prepared at different calcination temperatures.
SamplesCo3+(%)Co2+(%)Co3+/Co2+Oads(%)Olatt(%)Oads/Olatt
Co3O4-35053.2946.711.1441.4658.540.70
Co3O4-45051.6948.311.0636.4363.570.57
Co3O4-55050.2849.721.0133.0666.940.49
Table 3. XPS surface element analysis of catalysts prepared with different atmospheres.
Table 3. XPS surface element analysis of catalysts prepared with different atmospheres.
SamplesC(%)O(%)Co(%)Co3+/Co2+Oads/Olatt
Co3O4-35029.9047.3422.751.140.70
Co3O4-N2-O227.5550.0422.401.180.74
Co3O4-N264.0223.0612.920.961.92
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Jiang, S.; You, Z.; Tang, N. Effects of Calcination Temperature and Calcination Atmosphere on the Performance of Co3O4 Catalysts for the Catalytic Oxidation of Toluene. Processes 2023, 11, 2087. https://doi.org/10.3390/pr11072087

AMA Style

Jiang S, You Z, Tang N. Effects of Calcination Temperature and Calcination Atmosphere on the Performance of Co3O4 Catalysts for the Catalytic Oxidation of Toluene. Processes. 2023; 11(7):2087. https://doi.org/10.3390/pr11072087

Chicago/Turabian Style

Jiang, Sicheng, Zhimin You, and Ning Tang. 2023. "Effects of Calcination Temperature and Calcination Atmosphere on the Performance of Co3O4 Catalysts for the Catalytic Oxidation of Toluene" Processes 11, no. 7: 2087. https://doi.org/10.3390/pr11072087

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop